Gas Phase Hydrolysis of Formaldehyde To Form Methanediol: Impact

Apr 24, 2013 - Gas Phase Hydrolysis of Formaldehyde To Form Methanediol: Impact of Formic Acid Catalysis. Montu K. Hazra,*. ,†. Joseph S. Francisco,...
0 downloads 0 Views 800KB Size
Article pubs.acs.org/JPCA

Gas Phase Hydrolysis of Formaldehyde To Form Methanediol: Impact of Formic Acid Catalysis Montu K. Hazra,*,† Joseph S. Francisco,*,‡ and Amitabha Sinha*,§ †

Chemical Sciences Division, Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Kolkata-700064, India Department of Chemistry, Purdue University, West Lafayette, Indiana 47907-2084, United States § Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, California 92093-0314, United States ‡

S Supporting Information *

ABSTRACT: We find that formic acid (FA) is very effective at facilitating diol formation through its ability to reduce the barrier for the formaldehyde (HCHO) hydrolysis reaction. The rate limiting step in the mechanism involves the isomerization of a prereactive collision complex formed through either the HCHO···H2O + FA and/or HCHO + FA···H2O pathways. The present study finds that the effective barrier height, defined as the difference between the zero-point vibrational energy (ZPE) corrected energy of the transition state (TS) and the HCHO···H2O + FA and HCHO + FA···H2O starting reagents, are respectively only ∼1 and ∼4 kcal/mol. These barriers are substantially lower than the ∼17 kcal/mol barrier associated with the corresponding step in the hydrolysis of HCHO catalyzed by a single water molecule (HCHO + H2O + H2O). The significantly lower barrier heights for the formic acid catalyzed pathway reveal a new important role that organic acids play in the gas phase hydrolysis of atmospheric carbonyl compounds.

I. INTRODUCTION Formaldehyde is the simplest aldehyde and the most abundant carbonyl compound in our atmosphere.1 It is generated from the photochemical oxidation of hydrocarbons and the combustion of fossil fuel.1−3 Formaldehyde is also considered a major sink for the atmospheric oxidation of hydrocarbons and is an important source of atmospheric OH radicals and hydrogen molecules.2,4 In the liquid phase, the reaction between formaldehyde and water produces methanediol [CH2(OH)2], also known as the hydrated form of formaldehyde.5 Methanediol is of atmospheric interest because it is the smallest diol and diols have been implicated as potential seed molecules for aerosol growth.6 Methanediol is also an important molecule predicted to exist in interstellar medium, where it is believed to form in grain surface reactions triggered by the UV or cosmic ray processing of ice mantles.7,8 The gas phase hydrolysis of formaldehyde to form methanediol is known to be hindered by the presence of a substantial barrier.9−11 Thus, there is considerable current interest in understanding how diols form under atmospheric conditions. Recently, it has been shown that formic acid (FA) is able to lower the barrier for certain gas phase reactions.12 In this work we demonstrate for the first time the ability of formic acid, which is present in the atmosphere at significant trace levels,13 to effectively catalyze the hydrolysis of formaldehyde to form methanediol. For the purpose of direct comparison, we have also investigated the hydrolysis of formaldehyde catalyzed by © 2013 American Chemical Society

water molecules at the same level of theory. Our results suggest that the gas phase hydrolysis of formaldehyde catalyzed by formic acid is significantly more efficient compared to that involving catalysis by an equal number of water molecules. The present findings have major significance, as the results provide new insight into not only how methanediol might potentially be formed in the atmosphere but, by extension, the hydrolysis of other carbonyl compounds as well.

II. COMPUTATIONAL METHODS The Gaussian 03 and Gaussian 09 programs14 have been used to carry out all the quantum chemistry calculations presented here. The calculations have been performed using both the second order Møller−Plesset (MP2) perturbation theory and density functional theory (DFT). For the DFT calculations, the popular Becke’s three-parameter hybrid functional in conjunction with the Lee−Yang−Parr correlation (B3LYP) has been used.15 Transition states (TSs) for the water and FA assisted reactions have been located using the QST2/QST3 routines as implemented in the Gaussian 03 and Gaussian 09 programs. Furthermore, intrinsic reaction coordinate (IRC) Special Issue: Curt Wittig Festschrift Received: January 23, 2013 Revised: April 8, 2013 Published: April 24, 2013 11704

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

calculations were performed at the B3LYP/6-31+G(d,p) level to unambiguously verify that the transition states found connect with the desired reactants and products. The initial stationary point geometries were further optimized using the larger 6311++G(3df,3pd) basis set. Geometry optimization using the larger basis set is required to reduce basis set superposition error (BSSE), even though full (100%) counterpoise corrections often underestimate binding energies of dimeric complexes.16−18 In the case of FA catalyzed hydrolysis reaction we have also performed single point energy calculations at the CCSD(T) level in conjunction with 6-311++G(3df,3pd) and aug-cc-pVTZ basis sets to refine our energy estimates for the various stationary points. These calculations have been performed using optimized geometries obtained at the MP2/ 6-311++G(3df,3pd) level. The computed total electronic energies (Etotal) along with the zero-point energy (ZPE) corrected electronic energies [Etotal(ZPE)] of the monomers, complexes, and the transition states obtained at the DFT, MP2, and CCSD(T) levels are given in Supporting Information Table 1. Normal-mode vibrational frequency analyses have been performed for all the stationary points to verify that the stable minima have all positive vibrational frequencies and that the transition states have only one imaginary frequency (see Supporting Information Table 2). For all hydrolysis reactions, the B3LYP/6-311++G(3df,3pd) level predicted vibrational frequencies were used to estimate the respective ZPE corrections for the reactants, products, and TS.

Figure 1. Potential energy profile for the gas phase hydrolysis of HCHO involving a single water molecule. The energy profile has been calculated at the MP2/6-311++G(3df,3pd) level of theory with ZPE corrections.

this complex shows that our calculated values are similar to those obtained by Ramelot et al.20 including the orientation of the water and formaldehyde subunits forming the complex. The MP2/6-311++G(3df,3pd) level optimized geometries of the CH 2 (OH) 2 product and the TS associated with the HCHO···H2O → CH2(OH)2 unimolecular isomerization are also shown in Figure 1. All the stationary points on the potential energy surface have been shown to connect via an intrinsic reaction coordinate calculation. The binding energy of the starting HCHO···H2O complex at the MP2/6-311+ +G(3df,3pd) level is 3.7 kcal/mol and falls within the range (3.08−3.86 kcal/mol) predicted by Ramelot et al.20 From Figure 1 it is seen that the unimolecular isomerization of the prereactive complex to form methanediol, HCHO···H2O → CH2(OH)2, involves a four-member ring cyclic transition state with an activation energy of ∼39.2 kcal/mol measured relative to the ground state of the complex. From the figure we also see that the effective barrier for the reaction, measured relative to the free formaldehyde and a single water molecule, is 35.6 kcal/ mol at the MP2/6-311++G(3df,3pd) level. This value agrees well with the value of 36.3 kcal/mol computed by Kent et al. at the QCI(T)cc-pVTZ/QCI(T)cc-pVDZ level.7 Thus, the uncatalyzed hydrolysis of formaldehyde by a single water molecule has a barrier that is too large for efficient diol formation under typical atmospheric conditions. Studies show that the successive introduction of a second and third water molecule in the HCHO hydrolysis reaction progressively reduces the activation energy barrier for the unimolecular isomerization step.10,11 In order to facilitate a comparison with the results of HCHO hydrolysis catalyzed by formic acid, the hydrolysis reaction involving two water molecules is discussed next. As one of the waters can be replaced by formic acid, this allows an examination of the relative influence of the different catalyst molecules on the transition state. In enumerating the various possible geometries for the entrance channel complex involving two water molecules and HCHO, we have kept in mind the following basic principle regarding the hydrolysis reaction. The formation of methanediol from a single water and formaldehyde molecule basically involves two simultaneous steps: a hydrogen atom transfer and the formation of a carbon−oxygen bond between the water and formaldehyde units, as shown in Figure 1. Therefore, for the HCHO hydrolysis reaction involving two

III. RESULTS AND DISCUSSION The interaction between formaldehyde and water has been the subject of considerable prior interest.7,9−11,19−22 On the basis of these earlier studies, the uncatalyzed reaction between formaldehyde and a water molecule to produce methanediol can be written as follows: HCHO + H 2O ⇌ HCHO···H 2O → CH 2(OH)2

(1)

In the above reaction, HCHO···H2O is an entrance channel prereactive complex that isomerizes to methanediol via a fourmember ring cyclic TS. Among the several studies of the formaldehyde−water system we draw attention here to the study of Ramelot et al.,20 which appears to be the first study to explore the 1:1 formaldehyde−water molecule complex in detail using high level ab initio calculations at the CCSD and CISD levels in conjunction with basis sets that included both high angular momentum and diffuse functions. Their results suggested that the HCHO···H2O entrance channel complex has a planar ringlike structure with the O−H···OC and C−H···O groups forming angular hydrogen bonds. Later in 1996, the microwave spectrum of the formaldehyde−water complex was measured by Lovas et al.21 and their data supported a planar structure of the complex similar to that calculated by Ramelot et al.20 In the present work we have optimized the geometries of various monomers and associated complexes with larger basis sets and different methods to account for electron correlation, such as B3LYP/6-311++G(3df,3pd) and MP2/6311++G(3df,3pd) levels of theory. Figure 1 shows a summary diagram of the potential energy profile for the HCHO + H2O → CH2(OH)2 hydrolysis reaction with the corresponding data summarized in Supporting Information Tables 1, 3, and 4. From our calculations we find that the global minimum geometry of the formaldehyde−water prereactive complex is that represented by Structure-I in Figure 1. At the MP2/6-311+ +G(3df,3pd) level, an analysis of the geometrical parameters for 11705

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

water molecules, only geometries of the entrance channel complex that promote the above two simultaneous steps are expected to be effective in advancing the hydrolysis reaction (this is further illustrated in Figure 2). As the second water

Figure 2. Visualization of three possible structures of HCHO···H2O···H2O preassociation complex to understand the phenomena of active and passive catalysis in the formaldehyde hydrolysis reaction involving two water molecules.

Figure 3. Potential energy profile for the gas phase hydrolysis of HCHO involving two water molecules. The energy profile has been calculated at the MP2/6-311++G(3df,3pd) level of theory with ZPE corrections.

molecule is brought toward the HCHO···H2O complex, it can catalyze the hydrolysis reaction by acting either as a hydrogen atom acceptor (Structure-II) or as a hydrogen atom donor (Structure-III) or simultaneously as a hydrogen atom donor and acceptor (Structure-IV). Therefore in the entrance channel there are at least three different possible starting structures for the HCHO···(H2O)2 preassociation complex. However, it is found that only when the second water molecule acts as a bifunctional unit (i.e., simultaneously acting as both a hydrogen atom donor and acceptor), is the reaction barrier significantly reduced relative to that involving hydrolysis by a single water molecule.9 Thus, Structure-II and Structure-III cannot compete with Structure-IV with respect to their ability to reduce the reaction barrier.9 Having provided a rationale for selecting Structure-IV as the preferred entrance channel complex geometry, the hydrolysis reaction involving two water molecules can be written explicitly as

molecules is also not favored under atmospheric conditions, as the barrier still remains prohibitive. In fact, Wolfe et al.10 noted that even three water molecules are unable to reduce the barrier sufficiently to make the reaction feasible at atmospheric temperatures. Using formaldehyde with two waters as a template, we next explore the possibility of atmospheric species such as FA acting as a catalyst to promote this reaction. The hydrolysis of formaldehyde catalyzed by formic acid (FA) can be written in a manner analogous to that discussed above involving two water molecules: HCHO + H 2O + FA ⇌ HCHO···H 2O + FA ⇌ HCHO···H 2O...FA → CH 2(OH)2 ···FA ⇌ CH 2(OH)2 + FA

HCHO + 2H 2O ⇌ HCHO···H 2O + H 2O

or

⇌ HCHO···H 2O···H 2O → CH 2(OH)2 ···H 2O ⇌ CH 2(OH)2 + H 2O

HCHO + H 2O + FA ⇌ HCHO + FA···H 2O

(2)

⇌ HCHO···H 2O···FA → CH 2(OH)2 ···FA

or

⇌ CH 2(OH)2 + FA

HCHO + 2H 2O ⇌ HCHO + H 2O···H 2O

(5)

Like before, there are two main pathways for forming the HCHO···H2O···FA preassociation complex (eqs 4 and 5) which then undergoes unimolecular isomerization via an eightmember ring cyclic TS to form the CH2(OH)2···FA complex in the exit channel. To the best of our knowledge, there have been no previous studies reported in the literature investigating the reactions represented by eqs 4 and 5. In case of the HCHO···H2O + FA pathway (eq 4), we started with the optimized geometry of the formaldehyde−water complex (Structure-I in Figure 1) and then brought the formic acid moiety, at its isolated optimized geometry, toward the HCHO···H2O dimer to form various possible structures for the preassociation complex. The criterion used in forming these initial structures was to consider configurations in which formic acid is able to help the hydrolysis reaction in terms of simultaneously promoting carbon−oxygen bond formation and transfer a hydrogen atom between the water and formaldehyde moieties; these are the same criteria discussed above for the hydrolysis reaction involving two water molecules. Similarly, in the case of the HCHO + FA···H2O pathway (eq 5), we started

⇌ HCHO···H 2O···H 2O → CH 2(OH)2 ···H 2O ⇌ CH 2(OH)2 + H 2O

(4)

(3)

The above two equations represent the two pathways by which the hydrolysis reaction can occur. Basically, either a HCHO···H2O dimer complex can collide with a H2O molecule or HCHO can collide with the water dimer (H2O···H2O). In both cases the HCHO···H2O···H2O prereactive association complex is formed, which then isomerizes via a six-member ring cyclic TS to generate the exit channel complex between methanediol and a water molecule (CH2(OH)2···H2O). A schematic diagram of the potential energy profile for the HCHO hydrolysis reaction involving two water molecules, computed at the MP2/6-311++G(3df,3pd) level, is shown in Figure 3. The optimized geometries of the entrance and exit channel complexes as well as the TS are also shown in the figure. On the basis of the energetics summarized in Figure 3, it is clear that like the hydrolysis reaction involving a single water molecule, the hydrolysis of formaldehyde involving two water 11706

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

barrier height being less than ∼0.2 kcal/mol. This is not too surprising, since a small rotation of the HCHO subunit in the HCHO···H2O···FA preassociation complex can cause interconversion between the two structures. Consistent with these findings our calculations also show that the location of the transition state associated with the HCHO···H2O···FA → CH2(OH)2···FA isomerization reaction is independent of whether we take Structure-V or Structure-VI as the starting point. For the planar HCHO···H2O···FA complex (Structure-VI) the MP2/6-311++G(3df,3pd) level optimized geometries of the TS and the CH2(OH)2···FA exit channel complex are shown in Figure 5 along with the corresponding potential

with the optimized geometry of formaldehyde and then brought the FA···H2O dimer, with its geometry fixed at its optimized value, toward the HCHO moiety to form possible geometries of the HCHO···H2O···FA preassociation complex. Through this procedure a total of six preassociation complex geometries were generated for further consideration. The optimized structures of these six configurations are shown in Figure 4 along with their zero-point uncorrected relative energies

Figure 4. Six optimized geometries of the HCHO···H2O···FA entrance channel complex for the HCHO···H2O + FA and HCHO + FA···H2O pathways computed at the MP2/6-31G(d,p) level. Among the six geometries, the first three (Structure-V, Structure-VI, and Structure-VII) correspond to the HCHO···H2O + FA channel. Structure-V, StructureVI, Structure-VIII, Structure-IX, and Structure-X are associated with the HCHO + FA···H2O channel. Their zero-point uncorrected relative energies are also given.

Figure 5. Potential energy profile for the HCHO + H2O + FA → CH2(OH)2 + FA reaction associated with the starting planar geometry HCHO···H2O···FA preassociation complex. The energy profile has been calculated at the MP2/6-311++G(3df,3pd) level of theory with ZPE corrections.

computed at the MP2/6-31G(d,p) level. The first three geometries in Figure 4 are associated with the HCHO···H2O + FA channel, while the geometries represented by Structure-V, Structure-VI, Structure-VIII, Structure-IX, and Structure-X are associated with the HCHO + FA···H2O pathway. Hence, Structure-V and Structure-VI are common to both pathways. Further, as noted earlier, efficient methanediol formation through formaldehyde hydrolysis requires the formation of a carbon−oxygen bond between the formaldehyde and water moieties and the concomitant transfer of a hydrogen atom. From Figure 4 it is seen that among the structures shown, Structure-V and Structure-VI will be most effective in satisfying this requirement and reducing the reaction barrier. Basically, we find that with very little restructuring the preassociation complex geometries associated with Structure-V and StructureVI are able to undergo unimolecular isomerization to produce methanediol. From Figure 4, it is also seen that Structure-VII is very similar to Structure-VI. The only difference being that while Structure-VI is planar, in Structure-VII one of the hydrogen atoms of the water molecule forming the HCHO···H2O···FA complex is situated out of the plane formed by the heavy atoms in the complex. Thus, we focus only on Structure-V and Structure-VI, which are common to both the HCHO···H2O + FA and HCHO + FA···H2O channels. The fact that these two structures are true minima was confirmed by subsequent frequency calculations. Further, all calculations show that the interconversion of the HCHO···H2O···FA preassociation complex from its planar (Structure-VI) to the nonplanar geometry (Structure-V) is nearly a barrierless process with the

energy profile. The potential energy profile for the nonplanar preassociation complex (Structure-V) is given in the Supporting Information. When considering the planar preassociation complex (Structure-VI), the barrier height for the HCHO···H2O···FA → CH2(OH)2···FA unimolecular isomerization is 9.8 kcal/mol at the MP2/6-311++G(3df,3pd) level (see Figure 5). Further from the diagram it is seen that the energy difference between the HCHO···H2O + FA reactants and the TS associated with the HCHO···H2O···FA → CH2(OH)2···FA isomerization is only 0.1 kcal/mol with the TS being at higher energy. Therefore, unlike the hydrolysis reaction involving two water molecules, the formic acid catalyzed reaction is expected to be an energetically favorable mechanism for the formation of methanediol under atmospheric conditions. From Figure 5, it is seen that the energy of the TS for the HCHO···H2O···FA → CH2(OH)2···FA unimolecular isomerization is approximately 4−5 kcal/mol higher than that of the HCHO + FA···H2O reactants. Thus, the barrier associated with the HCHO + FA···H2O channel is computed to be higher than that for the HCHO···H2O + FA pathway. In order to improve our estimates of the reaction energetic, we have carried out single point energy calculations at the CCSD(T)/6-311++G(3df,3pd) and CCSD(T)/aug-cc-pVTZ levels using the MP2/6-311++G(3df,3pd) level geometries. From these two calculations we find that the energy difference between the HCHO···H2O + FA reactants and the TS increases respectively to 1.8 and 1.0 kcal/mol with the TS being at higher energy. The energies of the TS relative to the energy of 11707

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

Table 1. ZPE Corrected Relative Energies of the TS Associated with the HCHO···H2O···FA → CH2(OH)2···FA Unimolecular Isomerization Reaction for Both Planar and Nonplanar HCHO···H2O···FA Association Complex Starting Geometriesa level of theory

barrier height for the HCHO···H2O···FA → CH2(OH)2···FA unimolecular isomerization reaction, kcal/mol

B3LYP/6-311++G(3df,3pd) MP2/6-311++G(3df,3pd) CCSD(T)/6-311++G(3df,3pd) CCSD(T)/aug-cc-pVTZ

9.4 (9.1) 9.8 (10.1) 11.6 (11.8) 10.8 (11.1)

energy of TS relative to that of the HCHO···H2O + FA reactants, kcal/mol +0.61 +0.11 +1.76 +1.03

(+0.6) (+0.1) (+1.8) (+1.0)

energy of TS relative to that of the HCHO + FA···H2O reactants, kcal/mol +5.01 +4.42 +6.07 +5.39

(+5.0) (+4.4) (+6.1) (+5.4)

a

The energies are measured relative to the HCHO···H2O + FA and HCHO + FA···H2O reactants. The values in the parentheses are for the nonplanar HCHO···H2O···FA preassociation complex geometry.



HCHO···H2O + FA and HCHO + FA···H2O reactants (i.e., the reaction barrier height) determined using the various levels of theory are summarized in Table 1. The results shown suggest that the energy difference between the TS and the reactants is somewhat sensitive to the levels of calculation used. On the basis of the entries of the higher level calculations in Table 1, we believe that the energy of the TS is no more than 1−1.5 kcal/mol above the energy of the HCHO···H2O + FA reactants. Finally, we estimate the error associated with the calculated barrier height to be ±1.6 kcal/mol. This was determined by carrying out a series of calibration reactions calculations involving reactants and products reagents whose enthalpies are experimentally known and comparing with an analogous calculation where the reactant and product enthalpies were computed at the CCSD(T)/aug-cc-pVTZ level. These results are given in Supporting Information Table 5.

ASSOCIATED CONTENT

* Supporting Information S

Optimized geometries of reactants, products, and transition states in terms of their Z-matrices, their calculated total electronic energies, the zero-point vibrational energy corrected binding energies of all the hydrogen-bonded complexes at different levels of theory, the potential energy profile associated with the nonplanar HCHO···H2O···FA preassociation complex geometry as well as error estimates associated the computed energetics is also presented there. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; [email protected]; asinha@ ucsd.edu. Notes

The authors declare no competing financial interest.



IV. CONCLUSION The interaction between formaldehyde (HCHO) and water molecules to produce methanediol [CH 2(OH) 2] is of considerable interest because of its importance as a nucleophilic addition reaction in the fields of chemistry and biochemistry as well as in atmospheric studies. Even though formaldehyde and formic acid are present in our atmosphere in fairly significant abundance,13,23 the gas phase hydration of formaldehyde catalyzed by formic acid has yet to be considered. Results from the present study show that the barrier for the HCHO···H2O···FA → CH2(OH)2···FA unimolecular isomerization, which is the rate limiting step in formic acid catalysis, is substantially lower than that for the corresponding water catalyzed reaction.10 The findings from this work suggest that for catalysis by formic acid (FA), the transition state barrier is no more than 1.0−1.5 kcal/mol above the energy of the HCHO···H2O + FA reactants. The relatively low barrier associated with the HCHO···H2O + FA bimolecular encounter suggests that this reaction mechanism can lead to facile formation of methanediol under gas phase conditions. This mechanism is also expected to be favorable on cold water-rich interfaces such as ice and atmospheric aerosols where the HCHO···H2O complex can readily form and then subsequently interact with gas phase formic acid to form methanediol. Further, it is also quite likely that methanediol, which is one of several oxygen bearing species predicted to be present in interstellar medium,8 forms in such environments by the reaction mechanism discussed here. Finally, the present results suggest, more generally, that catalysis by organic acids may be an efficient common mechanism for the production of other atmospheric diols through the hydrolysis of appropriate carbonyl compounds in the gas phase.

ACKNOWLEDGMENTS A.S. thanks the National Science Foundation (U.S.) for the support of this work under Grant CHE-0642640. M.K.H. acknowledges the computational facilities of Saha Institute of Nuclear Physics, Kolkata, India, and the computational resource at the laboratory of Prof. T. Chakraborty (Department of Physical Chemistry, IACS-Kolkata, India).



REFERENCES

(1) Anderson, L. G.; Lanning, J. A.; Barrell, R.; Miyagishima, J.; Jones, R. H.; Wolfe, P. Sources and Sinks of Formaldehyde and Acetaldehyde: An Analysis of Denver’s Ambient Concentration Data. Atmos. Environ. 1996, 30, 2113−2123. (2) Parrish, D. D.; Ryerson, T. B.; Mellqvist, J.; Johansson, J.; Fried, A.; Richter, D.; Walega, J. G.; Washenfelder, R. A.; de Gouw, J. A.; Peischl, J.; et al. Primary and Secondary Sources of Formaldehyde in Urban Atmospheres: Houston Texas Region. Atmos. Chem. Phys. 2012, 12, 3273−3288. (3) (a) Altshuller, A. P. Production of Aldehydes as Primary Emissions and from Secondary Atmospheric Reactions of Alkenes and Alkanes during the Night and Early Morning Hours. Atmos. Environ. 1993, 27A, 21−32. (b) Finlayson-Pitts, B. J.; Pitts, J. N., Jr. Atmospheric Chemistry: Fundamental and Experimental Techniques; John Wiley & Sons: New York, 1986. (c) Seinfeld, J. H. Atmospheric Chemistry and Physics of Air Pollution; John Wiley & Sons: New York, 1986; (d) Pinto, J. P.; Gladstone, G. R.; Yung, Y. L. Photochemical Production of Formaldehyde in Earth’s Primitive Atmosphere. Science 1980, 210, 183−185. (e) Wofsy, S. C.; McConnell, J. C.; McElroy, M. B. Atmospheric CH4, CO, and CO2. J. Geophys. Res. 1972, 77, 4477− 4493. (4) (a) Jaeglé, L.; Jacob, D. J.; Wennberg, P. O.; Spivakovsky, C. M.; Hanisco, T. F.; Lanzendorf, E. J.; Hintsa, E. J.; Fahey, D. W.; Keim, E. R.; Proffitt, M. H.; et al. Observed OH and HO2 in the Upper 11708

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652. (16) Reimann, B.; Buchhold, K.; Barth, H.-D.; Brutschy, B.; Tarakeshwar, P.; Kim, K. S. Anisole-(H2O)n (n = 1−3) Complexes: An Experimental and Theoretical Investigation of the Modulation of Optimal Structures, Binding Energies, and Vibrational Spectra in Both the Ground and First Excited States. J. Chem. Phys. 2002, 117, 8805− 8822. (17) Alvarez-Idaboy, J. R.; Galano, A. Counterpoise Corrected Interaction Energies Are Not Systematically Better than Uncorrected Ones: Comparison with CCSD(T) CBS Extrapolated Values. Theor. Chem. Acc. 2010, 126, 75−85. (18) Galano, A.; Alvarez-Idaboy, J. R. A New Approach to Counterpoise Correction to BSSE. J. Comput. Chem. 2006, 27, 1203−1210. (19) (a) Morokuma, K. Molecular Orbital Studies of Hydrogen Bonds. III. CO···H−O Hydrogen Bond in H2CO···H2O and H2CO···2H2O. J. Chem. Phys. 1971, 55, 1236−1244. (b) Del Bene, J. E. Molecular Orbital Theory of the Hydrogen Bond. VI. The Effect of Hydrogen Bonding on the n → π* Transition in Dimers ROH···OCH2. J. Am. Chem. Soc. 1973, 95, 6517−6522. (c) Ahlström, M.; Jö nsson, B.; Karlströ m, G. Ab Initio Molecular Orbital Calculations on Hydrogen- and Non-Hydrogen-Bonded Complexes. H2CO·H2O and H2CO·H2S. Mol. Phys. 1979, 38, 1051−1059. (d) Williams, I. H.; Spangler, D.; Femec, D. A.; Maggiora, G. M.; Schowen, R. L. Theoretical Models for Transition-State Structure and Catalysis in Carbonyl Addition. J. Am. Chem. Soc. 1980, 102, 6619− 6621. (e) Williams, I. H.; Maggiora, G. M.; Schowen, R. L. Theoretical Models for Mechanism and Catalysis in Carbonyl Addition. J. Am. Chem. Soc. 1980, 102, 7831−7839. (f) Nelander, B. Infrared Spectrum of the Water Formaldehyde Complex in Solid Argon and Solid Nitrogen. J. Chem. Phys. 1980, 72, 77−84. (g) Butler, L. G.; Brown, T. L. Nuclear Quadrupole Coupling Constants and Hydrogen Bonding. Molecular Orbital Study of Oxygen-17 and Deuterium Field Gradients in Formaldehyde−Water Hydrogen Bonding. J. Am. Chem. Soc. 1981, 103, 6541−6549. (h) Williams, I. H.; Spangler, D.; Femec, D. A.; Maggiora, G. M.; Schowen, R. L. Theoretical Models for Solvation and Catalysis in Carbonyl Addition. J. Am. Chem. Soc. 1983, 105, 31−40. (i) Chin, S.; Ford, T. A. Ab Initio Molecular Orbital Calculations of the Infrared Spectra of Water Molecules in Some Intermolecular Complexes. J. Mol. Struct.: THEOCHEM 1985, 133, 193−209. (j) Kumpf, R. A.; Damewood, J. R., Jr. Interaction of Formaldehyde with Water. J. Phys. Chem. 1989, 93, 4478−4486. (k) Blair, J. T.; Westbrook, J. D.; Levy, R. M.; Krogh-Jespersen, K. Simple Models for Solvation Effects on Electronic Transition Energies: Formaldehyde and Water. Chem. Phys. Lett. 1989, 154, 531−535. (l) Vos, R. J.; Hendriks, R.; Van Duijneveldt, F. B. SCF, MP2, and CEPA-1 Calculations on the OH···O Hydrogen Bonded Complexes (H2O)2 and (H2O-H2CO). J. Comput. Chem. 1990, 11, 1−18. (m) Nelander, B. A Matrix Isolation Study of the Water−Formaldehyde Complex. The Far-Infrared Region. Chem. Phys. 1992, 159, 281−287. (n) Gerhards, M.; Kimpfel, B.; Pohl, M.; Schmitt, M.; Kleinermanns, K. Vibronic Spectroscopy of Jet-Cooled Clusters. J. Mol. Struct. 1992, 270, 301−324. (o) Francisco, J. S.; Williams, I. H. Reaction Pathways for Gas-Phase Hydrolysis of Formyl Compounds HXCO (X = H, F, and Cl). J. Am. Chem. Soc. 1993, 115, 3746−3751. (p) Fukunaga, H.; Morokuma, K. Cluster and Solution Simulation of Formaldehyde− Water Complexes and Solvent Effect on Formaldehyde 1(n, π*) Transition. J. Phys. Chem. 1993, 97, 59−69. (q) Ha, T.-K.; Makarewicz, J.; Bauder, A. Ab Initio Study of the Water−Formaldehyde Complex. J. Phys. Chem. 1993, 97, 11415−11419. (r) Dimitrova, Y.; Peyerimhoff, S. D. Theoretical Study of Hydrogen-Bonded Formaldehyde−Water Complexes. J. Phys. Chem. 1993, 97, 12731−12736. (20) Ramelot, T. A.; Hu, C.-H.; Fowler, J. E.; DeLeeuw, B. J.; Schaefer, H. F., III. Carbonyl−Water Hydrogen Bonding: The H2COH2O Prototype. J. Chem. Phys. 1994, 100, 4347−4354. (21) Lovas, F. J.; Lugez, C. L. The Microwave Spectrum and Structure of CH2O−H2O. J. Mol. Spectrosc. 1996, 179, 320−323.

Troposphere Suggest a Major Source from Convective Injection of Peroxides. Geophys. Res. Lett. 1997, 24, 3181−3184. (b) Warneck, P. Chemistry of the Natural Atmosphere, 2nd ed.; Academic Press: New York, 1999. (c) Novelli, P. C.; Lang, P. M.; Masarie, K. A.; Hurst, D. F.; Myers, R.; Elkins, J. W. Molecular Hydrogen in the Troposphere: Global Distribution and Budget. J. Geophys. Res. 1999, 104, 30427− 30444. (d) Sanderson, M. G.; Collins, W. J.; Derwent, R. G.; Johnson, C. E. Simulation of Global Hydrogen Levels Using a Lagrangian Three-Dimensional Model. J. Atmos. Chem. 2003, 46, 15−28. (5) Walker, F. J. Formaldehyde; Reinhold: New York, 1964. (6) Axson, J. L.; Takahashi, K.; De Hann, D. O.; Vaida, V. Gas-Phase Water-Mediated Equilibrium between Methylglyoxal and Its Geminal Diol. Proc. Natl Acad. Sci. U.S.A. 2010, 107, 6687−6692. (7) Kent, D. R., IV; Widicus, S. L.; Blake, G. A.; Goddard, W. A., III. A Theoretical Study of the Conversion of Gas Phase Methanediol to Formaldehyde. J. Chem. Phys. 2003, 119, 5117−5120. (8) Schutte, W. A.; Allamandola, L. J.; Sandford, S. A. Formaldehyde and Organic Molecule Production in Astrophysical Ices at Cryogenic Temperatures. Science 1993, 259, 1143−1145. (9) Williams, I. H. Theoretical Modelling of Specific Solvation Effects upon Carbonyl Addition. J. Am. Chem. Soc. 1987, 109, 6299−6307. (10) Wolfe, S.; Kim, C.-K.; Yang, K.; K. Weinberg, K.; Shi, Z. Hydration of the Carbonyl Group. A Theoretical Study of the Cooperative Mechanism. J. Am. Chem. Soc. 1995, 117, 4240−4260. (11) Böhm, S.; Antipova, D.; Kuthan, J. Methanediol Decomposition Mechanisms: A Study Considering Various ab Initio Approaches. Int. J. Quantum Chem. 1996, 58, 47−55. (12) (a) Hazra, M. K.; Sinha, A. Formic Acid Catalyzed Hydrolysis of SO3 in the Gas Phase: A Barrierless Mechanism for Sulfuric Acid Production of Potential Atmospheric Importance. J. Am. Chem. Soc. 2011, 133, 17444−17453. (b) Long, B.; Long, Z.-w.; Wang, Y.-b.; Tan, X.-f.; Han, Y.-h.; Long, C.-y.; Qin, S.-j.; Zhang, W.-j. Formic Acid Catalyzed Gas-Phase Reaction of H2O with SO3 and the Reverse Reaction: A Theoretical Study. ChemPhysChem 2011, 13, 323−329. (c) Hazra, M. K.; Chakraborty, T. Formamide Tautomerization: Catalytic Role of Formic Acid. J. Phys. Chem. A 2005, 109, 7621−7625. (d) Hazra, M. K.; Chakraborty, T. 2-Hydroxypyridine ↔ 2-Pyridone Tautomerization: Catalytic Influence of Formic Acid. J. Phys. Chem. A 2006, 110, 9130−9136. (e) Buszek, R. J.; Sinha, A.; Francisco, J. S. The Isomerization of Methoxy Radical: Intramolecular Hydrogen Atom Transfer Mediated through Acid Catalysis. J. Am. Chem. Soc. 2011, 133, 2013−2015. (f) da Silva, G. Carboxylic Acid Catalyzed Keto−Enol Tautomerizations in the Gas Phase. Angew. Chem., Ind. Ed. 2010, 49, 7523−7525. (g) Maeda, S.; Komagawa, S.; Uchiyama, M.; Morokuma, K. Finding Reaction Pathways for Multicomponent Reactions: The Passerini Reaction Is a Four Component Reaction. Angew. Chem., Ind. Ed. 2011, 50, 644−649. (13) (a) Shephard, M. W.; Goldman, A.; Clough, S. A.; Mlawer, E. J. Spectroscopic Improvements Providing Evidence of Formic Acid in AERI-LBLRTM Validation Spectra. J. Quant. Spectrosc. Radiat. Transfer 2003, 82, 383−390. (b) Grutter, M.; Glatthor, N.; Stiller, G. P.; Fischer, H.; Grabowski, U.; Höpfner, M.; Kellmann, S.; Linden, A.; von Clarmann, T. Global Distribution and Variability of Formic Acid As Observed by MIPAS-ENVISAT. J. Geophys. Res. 2010, 115, D10303. (c) Rinsland, C. P.; Mahieu, E.; Zander, R.; Goldman, A.; Wood, S.; Chiou, L. Free Tropospheric Measurements of Formic Acid (HCOOH) from Infrared Ground-Based Solar Absorption Spectra: Retrieval Approach, Evidence for a Seasonal Cycle, and Comparison with Model Calculations. J. Geophys. Res. 2004, 109, D18308. (14) (a) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; et al.et al. Gaussian 03, revision B.04; Gaussian, Inc.: Pittsburgh, PA, 2003. (b) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al.et al. Gaussian 09, revision B.01; Gaussian, Inc.: Wallingford, CT, 2010. (15) (a) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle− Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785−789. (b) Becke, A. D. Density11709

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710

The Journal of Physical Chemistry A

Article

(22) (a) Matubayasi, N.; Morooka, S.; Nakahara, M.; Takahashi, H. Chemical Equilibrium of Formaldehyde and Methanediol in Hot Water: Free-Energy Analysis of the Solvent Effect. J. Mol. Liq. 2007, 105, 58−63. (b) Mugnai, M.; Cardini, G.; Schettino, V.; Nielsen, C. J. Ab Initio Molecular Dynamics Study of Aqueous Formaldehyde and Methanediol. Mol. Phys. 2007, 134, 2203−2210. (23) Parrish, D. D.; Ryerson, T. B.; Mellqvist, J.; Johansson, J.; Fried, A.; Richter, D.; Walega, J. G.; Washenfelder, R. A; de Gouw, J. A.; Peischl, J.; et al. Primary and Secondary Sources of Formaldehyde in Urban Atmospheres: Houston Texas Region. Atmos. Chem. Phys. 2012, 12, 3273−3288.

11710

dx.doi.org/10.1021/jp4008043 | J. Phys. Chem. A 2013, 117, 11704−11710